Zirconium-Based Metal–Organic Frameworks for the Removal of

Nov 21, 2018 - toxins is critical for designing effective materials to remove these toxic compounds ... pies, such as a hemodialysis.6 Several porous ...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/JACS

Cite This: J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Zirconium-Based Metal−Organic Frameworks for the Removal of Protein-Bound Uremic Toxin from Human Serum Albumin Satoshi Kato,† Ken-ichi Otake,† Haoyuan Chen,‡ Isil Akpinar,† Cassandra T. Buru,† Timur Islamoglu,† Randall Q. Snurr,‡ and Omar K. Farha*,†,‡ †

Downloaded via WEBSTER UNIV on February 1, 2019 at 14:27:19 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Department of Chemistry and International Institute of Nanotechnology, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, United States ‡ Department of Chemical and Biological Engineering, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, United States S Supporting Information *

ABSTRACT: Uremic toxins often accumulate in patients with compromised kidney function, like those with chronic kidney disease (CKD), leading to major clinical complications including serious illness and death. Sufficient removal of these toxins from the blood increases the efficacy of hemodialysis, as well as the survival rate, in CKD patients. Understanding the interactions between an adsorbent and the uremic toxins is critical for designing effective materials to remove these toxic compounds. Herein, we study the adsorption behavior of the uremic toxins, p-cresyl sulfate, indoxyl sulfate, and hippuric acid, in a series of zirconium-based metal−organic frameworks (MOFs). The pyrene-based MOF, NU-1000, offers the highest toxin removal efficiency of all the MOFs in this study. Other Zr-based MOFs possessing comparable surface areas and pore sizes to NU-1000 while lacking an extended aromatic system have much lower toxin removal efficiency. From single-crystal X-ray diffraction analyses assisted by density functional theory calculations, we determined that the high adsorption capacity of NU-1000 can be attributed to the highly hydrophobic adsorption sites sandwiched by two pyrene linkers and the hydroxyls and water molecules on the Zr6 nodes, which are capable of hydrogen bonding with polar functional groups of guest molecules. Further, NU-1000 almost completely removes p-cresyl sulfate from human serum albumin, a protein that these uremic toxins bind to in the body. These results offer design principles for potential MOFs candidates for uremic toxin removal.



INTRODUCTION The accumulation of uremic retention solutes induces serious complications in patients with chronic kidney disease (CKD), a condition affecting approximately 14% of the general population.1 In high concentrations, uremic retention solutes are referred to as uremic toxins since they can become toxic to many physiological systems, causing issues like organ failure.2 In one example, p-cresol, one of the well-studied uremic toxins, synthesized by enteric bacteria disrupts a wide variety of biological activities, including endothelial dysfunction and the suppression of respiratory burst action in the blood. In addition, most of p-cresol undergoes conjugation with sulfates in the submucosal tissue via the action of sulfotransferases.3a Thus, p-cresyl sulfate is accumulated via organic anion transporters and enhances the production of reactive oxygen species in renal tubular cells resulting in cytotoxicity.3b Uremic toxins can be divided into several groups typically based on molecular weight and plasma protein-binding characteristics. Many small molecule uremic toxins, including p-cresyl sulfate, indoxyl sulfate, and hippuric acid, primarily bind to the transport protein human serum albumin (HSA) in human blood.4 These protein-bound uremic toxins (PBUTs) possess © XXXX American Chemical Society

an aromatic moiety and an ionic functional group, which allow for binding to several adsorption sites on HSA by electrostatic interaction and/or van der Waals forces.5 As a consequence, these interactions prevent the efficient removal of PBUTs through conventional extracorporeal renal replacement therapies, such as a hemodialysis.6 Several porous adsorbents targeting PBUTs have been reported, including activated carbon,7 zeolites,8−11 and composite membranes.12−14 However, these materials generally lack selectivity and/or efficiency;7,8,13 for example, activated carbons remove not only the protein-bound and solvated uremic toxins, but also other useful molecules such as pharmaceuticals because the adsorption performance relies solely on unselective van der Waals interactions.15 Furthermore, researchers are limited to indirect analysis methods, usually adsorption isotherms and kinetic tests or molecular simulations, to understand the adsorption behavior since activated carbon materials are amorphous. Crystalline adsorbents, on the other hand, would allow for direct structural characterization of adsorbates by Received: November 21, 2018

A

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society

Figure 1. Structure of (a) the Zr6 node core, (b) linkers for UiO-66, UiO-67 and UiO-NDC, (c) linker for MOF-808, (d) linker for NU-1000 and NU-901, (e) linker for NU-1010, (f) linker for PCN-608-OH, (g) linker for NU-1200, and (h) Zr6-based MOFs grouped by topology. Green = Zr, Red = O, Gray = C. In the structures, hydrogens are omitted for clarity.

synthesized following previously reported procedures except newly synthesized NU-1010, and details of all the syntheses are summarized in the Supporting Information (SI). The N2 isotherms and PXRD patterns of the MOFs are given in Figures S1 and S2 and Table S1. UiO-66, UiO-67, and UiONDC (Figure 1b) consist of 12-connected Zr6 nodes bridged by 1,4-benzenedicarboxylate (BDC) linkers, biphenyl-4,4dicarboxylate (BPDC) linkers, and 2,6-naphthalenedicarboxylate linkers (NDC), respectively.27 NU-1000,28 NU-1010, and PCN-608-OH29 consist of 8-connected Zr6 nodes bridged by tetratopic linkers 4,4′,4″,4″′-(pyrene-1,3,6,8-tetrayl)tetrabenzoate (TBAPy), 3,3′,5,5′-tetrakis(4-carboxyphenyl)1,1′-biphenyl (TCPB), 3,3′,5,5′-tetrakis(4-carboxyphenyl)4,4′-dihydroxybiphenyl (TCPB−OH), respectively, to give csq net topology structures. While NU-901 contains the same linker and node as in NU-1000, the torsion angles in the linker crystallize NU-901 in a different topology (scu) than NU1000.30 NU-1200 contains 8-connected Zr6 nodes bridged by 4′,4″-(2,4,6-trimethylbenzene-1,3,5-triyl)tribenzoate (TMTB) linkers to give the the net topology.31 Lastly, MOF-808 contains 6-connected Zr6 nodes bridged by tritopic trimesate linkers to afford spn topology.32 Knowing that the structures of uremic toxins contain an aromatic ring and an ionic functional group, the selection of these MOF materials attempted to exploit two types of interactions with the adsorbate: (1) hydrophobic interaction with the linker and (2) electrostatic interaction with the Zr6 node.33 If π−π interactions dominate, then UiO-NDC should exhibit higher toxin uptake than UiO-66 and UiO-67 due to its larger π-conjugated system. Additionally, if the hydroxyl ligands on the Zr6 node have strong interaction with uremic

single-crystal X-ray diffraction (SCXRD) analysis, assisting the determination of binding motifs and the design of superior adsorbent materials. Unlike other classes of porous, crystalline materials, metal−organic frameworks (MOFs) have exceptional tunability, allowing for systematic studies, as well as the incorporation of a host of functionalities. MOFs are composed of inorganic nodes bridged by multitopic organic linkers.16−18 Previously, several MOFs have been reported for the removal of water-soluble uremic toxin in an aqueous solution, such as a creatinine.18 Zr6-based MOFs in particular exhibit high chemical stability in aqueous media, and they have been tailored toward applications19 including, but not limited to, gas storage,20 gas separation,21 catalysis,22,23 drug delivery,24 and water remediation.25 Moreover, the crystalline nature of MOFs permits the unambiguous structural determination of the installed compounds through crystallographic techniques. Binding motifs of many oxyanions in MOFs have been observed and characterized by SCXRD analysis.26 However, there have been comparatively few investigations into the interactions between organic adsorbates and adsorbents, and even fewer studies into competitive adsorption. Here, we employ a series of Zr-based MOFs as PBUT adsorbents and probe the adsorption behavior of various uremic toxins in a MOF through a combination of direct and indirect analysis.



RESULTS AND DISCUSSION

Screening of Zr6 MOFs for p-Cresyl Sulfate Removal. To understand the relationships between the structural differences and adsorption behavior(s), nine Zr6-based MOFs, with varying linker structure, topology, and connectivity (Figure 1) were selected. The MOF materials were B

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society

on multiple structural factors, governing the adsorption behavior of p-cresyl sulfate in MOFs. Adsorption Site Analysis. To probe the adsorption mechanism of p-cresyl sulfate in a MOF, we chose NU-1000, which can readily adsorb the uremic toxin and be grown as single crystals suitable for XRD. We soaked single crystals of NU-1000 in potassium p-cresyl sulfate solution and attempted SCXRD analysis (Table S2). However, the aromatic region of p-cresyl sulfate is too diffuse and disordered to extract positions (Figures S4). Guided by the SCXRD data, the stable structure of p-cresyl sulfate was extracted using density functional theory (DFT, see Experimental Section). As displayed in Figure 3, p-cresyl sulfate is sited in two different locations at approximately the same occupancy within NU1000. In both cases, the average O−O distances from the hydroxyl/water group on Zr6-node to the sulfate moiety of pcresyl sulfate in the triangular pore and the hexagonal mesopore were determined to be 2.37 Å, a distance indicative of hydrogen bonding (Figures 3c and S5). The p-cresyl sulfate in the large hexagonal mesopore (Site 1) is perpendicular to the pyrene region and interacts with the phenyl substituent of pyrene linker at a distance reasonable for π−π interaction (4.24 Å).34 In contrast to Site 1, p-cresyl sulfate in the triangle small pore (Site 2) lies parallel to the pyrene region (Figure 3c) and interacts with two neighboring pyrene linkers (5.20 Å) (Figure S5c). Similarly, the SCXRD analyses (Table S2 and Figure S7) assisted by DFT calculation using indoxyl sulfate as an adsorbate reveal nearly identical hydrogen bonding and π− interactions between indoxyl sulfate and NU-1000 compared with p-cresyl sulfate in NU-1000 (Figure S8). These results indicate that not only electrostatic interactions and π−π interactions are factors in uremic toxin adsorption, but also maximizing these interactions through careful topology design can be important to capture p-cresyl sulfate. Specifically, the high p-cresyl sulfate uptake by NU-1000 is caused by highly hydrophobic adsorption sites sandwiched by two pyrene linkers and the nearby hydroxyl group on the Zr6 nodes capable of hydrogen bonding with the ionic functional groups of the adsorbates, all of which are carefully positioned (geometric parameters for π−π interactions are summarized in Table S3 and S4). Consequently, the UiO-type MOFs and

toxins, then MOF-808, possessing the most (12) hydroxyl groups per node, should have high uptake. Figure 2 displays the removal efficiency of 0.1 M potassium p-cresyl sulfate solution, a concentration which simulates the

Figure 2. Removal efficiencies of 1.5 mg of MOF in 0.1 mM aqueous potassium p-cresyl sulfate solution (pH ≈ 6.5, orange: 12 connected MOF, blue: 8 connected MOF, red: 6 connected MOF).

blood of CKD patients.3,38 The corresponding molar uptakes are described in Figure S3. Under these conditions, NU-1000 offered the highest uptake, removing 94% of the p-cresyl sulfate in the solution. Contrary to the anticipated results, the uptake of UiO-NDC (3.3%) was found to be nearly identical to that of UiO-66 (2.1%) and UiO-67 (4.7%). Interestingly, the uptake of the 6-connected MOF-808 (6.2%) was significantly lower than the uptakes of some 8-connected MOFs (NU-1000, NU901, NU-1010). Curiously, PCN-608-OH and NU-1200 exhibited significantly lower uptake than NU-1000 although these MOFs have high surface area, same csq topology and/or connectivity. Further, NU-901 containing the pyrene-based linker displayed the second highest uptake (79%), despite NU901 having the smallest pores/channels (Table S1, see SI). Slight variation of torsion angles in the linker structure could perturb the MOF−toxin interaction enough to disrupt analyte adsorption. These results hint at a complex mechanism, relying

Figure 3. (a) The crystal structure of NU-1000 viewed down the c-axis and (b) an orthogonal view prior to p-cresyl sulfate exposure (each adsorption site is depicted by a colored oval). (c) Optimized geometry of p-cresyl sulfate-pyrene and Zr6 node domains after p-cresyl sulfate adsorption. For clarity, only one orientation of p-cresyl sulfate is extracted at each site, and potassium counterions and hydrogens are omitted. C

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society MOF-808 with insufficient π-conjugated system and/or hydroxyl groups on Zr6 nodes offered low p-cresyl sulfate uptake. Although NU-1200 has large pore sizes allowing for rapid diffusion, steric repulsion can exist between p-cresyl sulfate and the methyl group on the linker, preventing adsorption in NU-1200. To understand how the structure of MOF linker affects uptake, we compared NU-1000, NU-1010, and PCN-608-OH. The three MOFs have the same (4,8)connected network with csq net and similar porosity, but only NU-1000 exhibited high removal efficiency. NU-1010 has smaller π-conjugated linkers, and its uptake efficiency is only 29% since the π−π interactions of p-cresyl with biphenyl at Site 2 are weaker than those with pyrene. Moreover, the linker in PCN-608-OH is structurally similar to the linker in NU-1010 except the 4,4′-positions of the biphenyl are substituted with a hydroxyl group. As a result, the uptake decreased to 4.6% likely owing to the steric hindrance caused by the hydroxyl groups (Figure S9). Kinetic and Adsorption Isotherm Studies. To understand the effect of uremic toxins’ structures on adsorption behavior, we compared the adsorption of two more uremic toxins as adsorbates, indoxyl sulfate (as K+ salt) and hippuric acid (Figure 4a). These uremic toxins are also commonly

isotherms of these uremic toxins in NU-1000 are given in Figure 5. The observed Type I isotherms show similar trends to previously reported adsorption isotherms of uremic toxins in zeolites11 and activated carbon.12 After adsorption of p-cresyl sulfate, the DFT-calculated pore size of the large hexagonal mesopore derived from the N2 isotherm at 77 K slightly shifted from 29.5 to 27.3 Å and the BET area decreased from 2140 ± 5 m2/g to 1860 ± 15 m2/g (Figure S10). This decrease is attributed to the adsorption of p-cresyl sulfate in NU-1000 (Figure S11). Importantly, the crystallinity of the material was maintained after adsorption (Figure S12) so that the material could be easily recycled (Figure S13). A similar porosity reduction was observed after the adsorption of indoxyl sulfate and hippuric acid (Figure S10). To further elucidate the effect of the structure of uremic toxins on adsorption behavior, adsorption isotherms in NU-1000 at 303 and 310 K were collected (Figure 5). Adsorption equilibrium information was calculated by the Langmuir equation,35 and the Langmuir parameters Langmuir adsorption equilibrium constant (KL) and Langmuir adsorption capacity (Qmax)were determined from the transformed isotherms data (Table 1). The KL value for pcresyl sulfate slightly decreased from 10.6 to 1.8 mM−1 at 297 and 303 K, respectively. At 310 K, the p-cresyl sulfate adsorption isotherm is nearly linear, best fitting the Freundlich model (R2 = 0.992, Figure S14),36 indicating that π−π interactions between NU-1000 and p-cresyl sulfate are weakened at elevated temperature. In contrast to the KL values of p-cresyl sulfate, the KL value for indoxyl sulfate remained fairly constant, ranging from 170 to 194 mM−1 upon increasing temperature. This is likely a consequence of a strong π−π interaction between indole moiety and pyrene linker of NU-1000; this π−π interaction was investigated through SCXRD analysis (Figure S8). The KL value for hippuric acid increased drastically from 197 to 703 mM−1 upon increasing temperature likely due to the exchange of hydroxyl groups on Zr6-node with hippuric acid promoted by thermal energy. Such effect has also been observed with other molecules containing carboxylic acid functional groups through solvent-assisted ligand incorporation (SALI) reactions (Figure S15).37 This exothermal effect also supports the observation that hippuric acid could be removed efficiently by MOFs other than NU1000 (Figure S16). The lowest energy configuration of hippuric acid in NU-1000 was calculated by DFT assuming that the carboxylic acid coordinates to the Zr6 node (Figure S17). The calculation indicated that hippuric acid can interact with pyrene linker in a similar way to p-cresyl sulfate and indoxyl sulfate when bound to the node. These analyses of adsorption isotherm affirm that Zr6-based MOFs can efficiently

Figure 4. (a) The structures of the uremic toxins used in this study and (b) the removal efficiency of 6 mg of NU-1000 in 0.1 mM aqueous solutions of uremic toxins as a function of time at 297 K (pH 6.5).

found bound to proteins.4 Figure 4b shows the removal efficiency of 0.1 mM uremic toxins by NU-1000. Within only 1 min, NU-1000 removed more than 70% of p-cresyl sulfate in solution. This rapid uptake can be attributed to the large aperture and pore volume of NU-1000 facilitating facile diffusion of p-cresyl sulfate. In addition, the uptakes of indoxyl sulfate and hippuric acid reached more than 98% in 1 min, suggesting stronger interactions between these two adsorbates and NU-1000 compared with p-cresyl sulfate. Adsorption

Figure 5. Adsorption isotherms of the uremic toxins on NU-1000 at 297, 303, and 310 K: (a) potassium p-cresyl sulfate, (b) potassium indoxyl sulfate, and (c) hippuric acid. The fitting of the isotherms generated from the Langmuir parameters in Table 1. The Langmuir plots are shown in Figure S14. D

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society Table 1. Langmuir Parameters for Uremic Toxins in NU-1000 adsorbate

temperature (K)

KL (mM−1)a

potassium p-cresyl sulfate

297 303 310 297 303 310 297 303 310

10.6 1.8

166 440

1.24 3.30

194 170 189 197 506 703

156 193 254 189 182 199

1.08 1.30 1.71 1.97 1.82 1.88

potassium indoxyl sulfate

hippuric acid

Qmax (mg g−1

b NU‑1000)

Qmax (mol mol−1

c NU‑1000)

a

Langmuir adsorption equilibrium constant. bAdsorption capacity on a mass basis. cAdsorption capacity on a mole basis. KL and Qmax were calculated by the Langmuir equation.

adsorb uremic toxins containing sulfate or carboxylic acid moieties. Predicted Removal of p-Cresyl Sulfate from Human Serum Albumin (HSA). In the human plasma of CKD patients, about 80% of p-cresyl sulfate is bound to human serum albumin (HSA, Figure 6a).3,38 Under our conditions, we measured a comparable average of 82 ± 5% of p-cresyl sulfate adsorbed on HSA at the initial concentration (Ci) range of 0.01 to 0.1 mM (Figure S18). We investigated the removal of p-cresyl sulfate from HSA by NU-1000. First, to understand the adsorption behavior of p-cresyl sulfate on HSA, adsorption isotherms were constructed (Figure 6b). We used the Freundlich model to fit for p-cresyl sulfate adsorption on HSA at 310 K. From the Freundlich parameters (Table S6), NU-1000 displayed higher gravimetric uptake of p-cresyl sulfate on HSA and larger adsorption constant, KF, than that of HSA. We also calculated molar uptake on each adsorbent (Figure S19). Additionally, NU-1000 maintains p-cresyl sulfate affinity in the presence of competing salt ions in aqueous solution of 0.2 M NaCl (Figure S20). With these results in mind, we hypothesized that the removal efficiency of p-cresyl sulfate from HSA could be estimated by comparing the predicted uptakes given by Freundlich parameters before the competitive adsorption in the presence of salt. From the derived Freundlich parameters, the predicted uptake for each adsorbent, Qe, can be described by eq 1. Q e(mg g −1) = KF × C1/ n

Q m,NU − 1000(%) =

Q e,NU − 1000 × MNU − 1000 Q e,NU − 1000 × MNU − 1000 + Q e,HSA × MHSA

× 100

(2) Q m,HSA(%) =

Q e,HSA × MHSA Q e,NU − 1000 × MNU − 1000 + Q e,HSA × MHSA

× 100

(3)

where M is the amount of adsorbent. In Figure 7, dotted line shows the predicted removal fraction of p-cresyl sulfate on each adsorbent. Herein, MHSA = 50 mg, MNU‑1000 = 2.5, 10, 20, 30 mg, and C = 0.001 to 0.1 mM. The calculation predicted that 93% of p-cresyl sulfate could be removed by adding 20 mg of NU-1000. To evaluate the reliability of the predicted p-cresyl sulfate removal from HSA, 20 μg of p-cresyl sulfate was mixed with 50 mg of HSA in 1.0 mL of an aqueous solution of 0.2 M NaCl at 310 K. After 24 h, various amounts of NU-1000 were added to each solution, and the solutions were kept for an additional 24 h at 310 K. After this competitive adsorption, the amounts of unbound and HSA-bound p-cresyl sulfate were measured by HPLC (details in Experimental Section). As shown in Figure 7, the fraction of HSA-bound and unbound pcresyl sulfate decreased as of the amount of added NU-1000 increased. After adding 20 mg of NU-1000, about 93% of pcresyl sulfate in the solution was removed, consistent with the predicted removal efficiency (93%). As shown in Table S7, the predicted values correlate well with the experimental competitive adsorption. Given the results, we now have a tool to estimate the removal efficiency of uremic toxins by only performing adsorption isotherms for the adsorbents.

(1)

where C is the residual concentration of p-cresyl sulfate, and 1/ n is the degree of nonlinearity (Table S6). Furthermore, the predicted adsorbed amount (Qm) could be calculated by eqs 2 and 3,

Figure 7. Adsorption study of p-cresyl sulfate with NU-1000 and human serum albumin (HSA) at 310 K in an aqueous solution of 0.2 M NaCl. Dotted line is the predicted fraction of adsorbed potassium p-cresyl sulfate on each adsorbent. Squares are the values determined experimentally (pH 6.5).

Figure 6. (a) The structure of HSA, (b) the adsorption isotherm of pcresyl sulfate on NU-1000 and HSA. C = the residual concentration of potassium p-cresyl sulfate at 310 K (pH 6.5). E

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society



Single-crystal X-ray diffraction measurements were performed on a Bruker Kappa APEX II CCD equipped with a Mo Kα (λ = 0.710 73 Å) IμS microsource with MX optics. A single crystal was mounted on MicroMesh (MiTeGen) with paratone oil. The structure was solved by direct methods (SHELXT-2014/5)39 and refined by full-matrix least-squares refinements on F2 (SHELXL-2017/1)40 using the Yadokari-XG software package.41 The disordered noncoordinated solvents were removed using the PLATON SQUEEZE program.42 The sulfate site locations and occupancies were determined by structural refinement. Refinement results are summarized in Table S1. All DFT calculations were carried out in the Gaussian 16 package43 using the M06-L density functional and def2-SVP basis set. The SDD effective core potential was applied to the Zr atoms. The DFT-D3 dispersion correction with zero damping was applied. Screening of p-Cresyl Sulfate Removal. In a typical experiment, 1.5 mg of MOF sample was immersed in 2.5 mL of a 0.1 mM aqueous potassium p-cresyl sulfate solution in a 1.5-dram glass vial. The solutions were placed in the oven at 24 °C for 24 h to obtain saturation uptake. After 24 h, the samples were filtered off through a syringe filter (PVDF membrane, 0.45 μm, Sartorius Minisart Syringe Filter). The initial and equilibrium concentrations of p-cresyl sulfate in each solution were measured using an Agilent HPLC 1100 system. The removal efficiency of p-cresyl sulfate was calculated by using eq 4:

CONCLUSIONS Adsorption of p-cresyl sulfate has been explored by using nine Zr6-based MOFs. NU-1000 was found to exhibit the highest capacity of all MOFs investigated here. The combination of SCXRD analyses and DFT calculations suggested that both electrostatic interaction with hydroxyl groups on the Zr6 nodes and π−π interaction with pyrene-based linkers play important roles in the adsorption of p-cresyl sulfate. Furthermore, NU1000 has two different adsorption sites in the large hexagonal mesopore (Site 1) and the small triangle micropore (Site 2). Compared with MOFs of the same topology, NU-1000 has superior performance in uremic toxin adsorption since the pcresyl sulfate can interact with the highly hydrophobic adsorption sites sandwiched by two pyrene linkers and the nearby hydroxyl group on the Zr6 nodes capable of hydrogen bonding with the ionic functional groups of the adsorbates. The Langmuir or Freundlich model was used to fit adsorption isotherms of different uremic toxins in the MOFs to calculate the adsorption parameters for obtaining kinetic parameters. The calculated adsorption parameters were found to agree with the results obtained through SCXRD analysis. Finally, we described a method to predict the uremic toxin removal from HSA by MOFs. The strong interactions, encompassing the π−π interaction with pyrene-based linker and hydrogen bonding to hydroxyl group on Zr6-nodes, make NU-1000 an efficient adsorbent for the removal of PBUTs. This study highlights important motifs which allow for high adsorption of PBUTs and paves the way for similar designer materials.



removal (%) =

Ci − Cf × 100 Ci

(4)

where Ci = the initial concentration and Cf = the final concentration. Adsorption Site Analysis. Single crystals of NU-1000 were soaked in an aqueous potassium p-cresyl sulfate solution (5.0 mM) at room temperature for 24 h and examined through SCXRD analysis. To determine the location of the adsorbed sulfates positions in the SCXRD analyses, the NU-1000 framework atoms were first located and refined, after which the residual electron densities were calculated. As shown in the F0−Fc difference Fourier maps (Figures S4 and S6), the site positions of “SO3” of the sulfates were clearly observed. However, the rest of the structure was difficult to model and refine due to the highly disordered nature of the aromatic ring compared to the high symmetry of the rest of the crystal. Therefore, to further address the location of theses guest molecules, DFT calculations were conducted. In the second step, based on the SCXRD data, geometry optimizations of p-cresyl and indoxyl sulfates were carried out in the Gaussian 16 package43 using the M06-L density functional and def2SVP basis set. The positions of the framework atoms were taken from the experimental crystal structures and fixed during the optimizations. Proper truncations of the distant atoms were made to reduce the system size while keeping the entire system charge neutral substituting the protons for the potassium cations. The proton topologies of the Zr6 nodes were built according to a previous study.18 For p-cresyl sulfate, the positions of the sulfur atoms were also fixed to the experimental data obtained from X-ray single crystal experiment while the other atoms were added and allowed to move. In the case of hippuric acid, the structure was built by replacing an adjacent pair of OH and OH2 groups on the node with the deprotonated hippuric acid through the coordination between two Zr atoms and the carboxylate group. The geometry optimizations were performed in the same way as described above in which the entire hippuric acid was not fixed and allowed to relax. Coordinates of all the optimized structures are available in the SI. Kinetic Isotherm Studies. To investigate the uptake of uremic toxins as a function of time, kinetic studies were performed by exposing 6 mg of NU-1000 to 10 mL of a 0.1 mM aqueous uremic toxin solutions, potassium p-cresyl sulfate, potassium indoxyl sulfate, and hippuric acid in a 6-dram vials. The solution was placed in the oven at 24 °C oven. At predetermined times, 1.5 mL aliquots were taken by filtering with a 0.45 μm PVDF syringe filter. The removal efficiency as a function of time was measured by HPLC system and calculated by eq 4. Adsorption Isotherm Studies Using NU-1000. Adsorption isotherms were constructed by exposing 1.5 mg of NU-1000 to 2.5

EXPERIMENTAL SECTION

Materials and Methods. All chemicals were used as-received from the supplier. In these experiments, Milli-Q water (Millipore) was used. Potassium p-cresyl sulfate was purchased from Enamine, potassium indoxyl sulfate was purchased from Alfa Aesar, and hippuric acid was purchased from Sigma-Aldrich. HPLC-GC grade acetonitrile (Sigma-Aldrich) and potassium dihydrogen phosphate (KH2PO4) (Strem chemicals) were used for all HPLC experiments as a mobile phase. UiO-66, UiO-67, UiO-NDC, PCN-608-OH, NU-901, NU-1000, NU-1200, and MOF-808 were prepared according to literature procedures (see SI).27−32 PXRD patterns were collected on an ATX-G (Rigaku) instrument equipped with an 18 kW copper rotating anode X-ray source. Roughly 3 mg of sample was loaded onto a sample holder and mounted on the instrument. Samples were recorded from 2° < θ < 20°at a scan rate of 2°/min and a step size of 0.05°. N2 sorption isotherms were collected on a Micromeritics Tristar II 3020 instrument at 77 K. Prior to the measurement, the samples were activated on a Smart VacPrep port by heating at desired temperature under vacuum overnight. Density functional theory (DFT)-calculated pore size distributions were calculated using a carbon slit-pore model with a kernel, based on molecular statistical approach. At least 20 mg of material was used for sorption measurements. High performance liquid chromatography (HPLC) measurements were performed by an Agilent HPLC 1100 system equipped with a binary pump, a column thermostat, an auto sampler, and a diode-array absorbance detector (DAD), using an Ascentic C18 HPLC column (5 μm particle, 150 × 4.6 mm2 I.D.). A reverse phase HPLC assay of pcresyl sulfate was carried out using an isocratic elution with a flow rate of 1.0 mL/min, a column temperature of 40 °C, a mobile phase of acetonitrile and water (35:65 v/v %) and a detection wavelength of 230 nm. Assays of indoxyl sulfate and hippuric acid were performed using a mobile phase of 0.02 mM KH2PO4 and 0.002 mM 1pentasulfonic acid (pH 4.50) and a detection wavelength at 270 and 230 nm, respectively. For all experiments, 100 μL sample volumes were injected into the chromatographic system. F

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society ORCID

mL of an aqueous water solution (or 0.2 M NaCl solution) with varying concentrations of uremic toxins, in 1.5-dram vials. The solutions were placed in the oven at the desired temperature (297, 303, or 310 K) oven for 24 h to reach saturation uptake. After 24 h, the sample was filtered with a 0.45 μm PVDF filter syringe. The initial and equilibrium concentrations were measured by using an Agilent HPLC 1100, and the amount of uremic toxins uptake Q in mg of uremic toxins per gram of MOF were determined at each point according to eq 5,

Q = (Ci − Cf ) ×

V m

Satoshi Kato: 0000-0002-3654-2122 Ken-ichi Otake: 0000-0002-7904-5003 Haoyuan Chen: 0000-0002-8634-4028 Cassandra T. Buru: 0000-0001-6142-8252 Timur Islamoglu: 0000-0003-3688-9158 Randall Q. Snurr: 0000-0003-2925-9246 Omar K. Farha: 0000-0002-9904-9845 Notes

(5)

The authors declare the following competing financial interest(s): O.K.F. and R.Q.S. have financial interest in NuMat Technologies, a startup company that is seeking to commercialize MOFs.

where V = volume of solution exposed to NU-1000 (2.5 mL here) and m = mass of NU-1000 in grams (1.5 mg here). All adsorption isotherms of uremic toxins adsorption on MOFs were described by either the Langmuir model or the Freundlich model, which are given in eqs 2 and 6. The Langmuir equation gives the Langmuir equilibrium constant (KL) and adsorption capacity (Qe) for each uremic toxin.

Qe =



ACKNOWLEDGMENTS Authors gratefully acknowledge the support of the Nanoporous Materials Genome Center, funded by the U.S. DOE, Office of Science, Basic Energy Sciences Program (Award DE-FG0217ER16362). We gratefully acknowledge the support of Asahi Kasei Corporation. We thank Arabela Grigorescu, James Casey and Ronald Soriano of the Northwestern Keck Biophysics Facility for assistance with HPLC instrument, data collection. We acknowledge the use of the resources of the Keck Biophysics Facility, supported by the NCI CCSG P30 CA060553 grant awarded to the Robert H Lurie Comprehensive Cancer Center of Northwestern University. This work made use of the IMSERC at Northwestern University, which has received support from the NSF (CHE-1048773 and DMR0521267); SHyNE Resource (NSF NNCI-1542205); the State of Illinois and International Institute of Nanotechnology.

KL × Q max × C 1 + KL × C

(6)

−1

where Qe (mg g ) = the amount adsorbed per unit amount of adsorbent at equilibrium, Qmax = the maximum uptake, and C = the residual concentration of adsorbate. Adsorption Isotherm Studies Using HSA. Adsorption isotherm studies were performed by exposing 50 mg of HSA to 1.0 mL of an aqueous solution of 0.2 M NaCl, designated concentration of potassium p-cresyl sulfate, in 1.5-dram vials. The solution was placed in the oven at 310 K oven for 24 h to reach saturation uptake. After 24 h, the sample was filtrated and centrifuged at 1000 RCF for 15 min to allow MOF to be settled by ultrafiltration device (ultracel PL membrane, 30 kDa, Millipore Sigma). The initial and equilibrium concentrations were measured by using an Agilent HPLC 1100, and the amount of uremic toxins uptake Q in mg of uremic toxins per gram of HSA was determined at each point according to eq 5. p-Cresyl Sulfate Removal from HSA (Competitive Adsorption). Competitive adsorption studies were performed by adding 50 mg of HSA to 1 mL of 0.2 M NaCl containing 20 μg of potassium pcresyl sulfate in 2-dram vials. The solution was placed in the oven at 310 K for 24 h ensure to reach saturation uptake, and designated amount of NU-1000 was added to this solution and kept in the oven. After 24 h, the sample was divided into two, and one-half was filtered and centrifuged at 1000 RCF for 15 min by ultrafiltration device (ultracel PL membrane, 30 kDa, Millipore Sigma). The other half was heated at 100 °C for 5 min and centrifuged at 3200 RCF for 5 min. The aliquot was filtered with a 0.45 μm PVDF filter syringe. The residual concentrations were measured by HPLC and removal efficiency as a function of time was calculated by eq 4.





(1) National Institute of Diabetes and Digestive and Kidney Diseases: Kidney Diseases Statics for the United States. https://www. niddk.nih.gov/health-information/health-statistics/kidney-disease (accessed Oct 8, 2018). (2) (a) Sun, C. Y.; Chang, S. C.; Wu, M. S. Uremic Toxins Induce Kidney Fibrosis by Activating intrarenal Renin-Angiotensin-Aldosterone System Associated Epithelial-to-Mesenchymal Transition. PLoS One 2012, 7, No. e34026. (b) Gryp, T.; Vanholder, R.; Vaneechoutte, M.; Glorieux, G. p-Cresyl sulfate. Toxins 2017, 9, 52. (3) (a) Martinez, W. A.; Recht, S. N.; Hostetter, H. T.; Meyer, W. T. Removal of P-cresol sulfate by Hemodyalisis. J. Am. Soc. Nephrol. 2005, 16, 3430−3436. (b) Watanabe, H.; Miyamoto, Y.; Honda, D.; Tanaka, H.; Wu, Q.; Endo, M.; Noguchi, T.; Kadowaki, D.; Ishima, Y.; Kotani, S.; Nakajima, M.; Kataoka, K.; Kim-Mitsuyama, S.; Tanaka, M.; Fukagawa, M.; Otagiri, M.; Maruyama, T. p-cresyl sulfate causes renal tubular cell damage by inducing oxidative stress by activation of NADPH oxidase. Kidney Int. 2013, 83, 582−592. (4) Vanholder, R.; De Smet, R.; Glorieux, G.; Argiles, A.; Baurmeister, U.; Brunet, P.; Clark, W.; Cohen, G.; De Deyn, P. P.; Deppisch, R.; Descamps-Latscha, B.; Henle, T.; Jorres, A.; Lemke, H. D.; Massy, Z. A.; Passlick-Deetjen, J.; Rodriguez, M.; Stegmayr, B.; Stenvinkel, P.; Tetta, C.; Wanner, C.; Zidek, W. Review on uremic toxins: Classification, concentration, and interindividual variability. Kidney Int. 2003, 63, 1934−1943. (5) (a) Yu, S.; Schuchardt, M.; Tölle, M.; van der Giet, M.; Zidek, W.; Dzubiella, J.; Ballauff, M. Interaction of human serum albumin with uremic toxins: a thermodynamic study. RSC Adv. 2017, 7, 27913−27922. (b) Berge-Lefranc, D.; Chaspoul, F.; Cerini, C.; Brunet, P.; Gallice, P. Thermodynamic study of indoxylsulfate interaction with human serum albumin and competitive binding with p-cresylsulfate. J. Therm. Anal. Calorim. 2014, 115, 2021−2026. (c) Ghuman, J.; Zunszain, P. A.; Petitpas, I.; Bhattacharya, A. A.;

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.8b12525. MOF syntheses, N2 isotherms, PXRD patterns, crystallographic data, molar removal efficiency of uremic toxins data, NMR spectra, recyclability test, and Freundlich parameters (PDF) Optimized geometry data and crystallographic data structures (CIF) Optimized geometry data and crystallographic data structures (CIF)



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*[email protected] G

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society

in a metal-organic microporous material. Nature 2005, 436, 238−241. (b) Li, B.; Wen, H.-M.; Zhou, W.; Chen, B. Porous Metal-Organic Frameworks for Gas Storage and Separation: What, How, and Why? J. Phys. Chem. Lett. 2014, 5, 3468−3479. (21) Bachman, J. E.; Smith, Z. P.; Li, T.; Xu, T.; Long, J. R. Enhanced ethylene separation and plasticization resistance in polymer membranes incorporating metal-organic framework nanocrystals. Nat. Mater. 2016, 15, 845−849. (22) (a) Peters, A. W.; Li, Z.; Farha, O. K.; Hupp, J. T. Atomically Precise Growth of Catalytically Active Cobalt Sulfide on Flat Surfaces and within a Metal-Organic Framework via Atomic Layer Deposition. ACS Nano 2015, 9 (8), 8484−8490. (b) Otake, K. I.; Cui, Y.; Buru, C. T.; Li, Z.; Hupp, J. T.; Farha, O. K. Single-Atom-Based Vanadium Oxide Catalysts Supported on Metal-Organic Frameworks: Selective Alcohol Oxidation and Structure-Activity Relationship. J. Am. Chem. Soc. 2018, 140, 8652−8656. (23) (a) Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-based metal-organic frameworks: design. synthesis, and applications. Chem. Soc. Rev. 2016, 45, 2327−2367. (b) Bernales, V.; Ortuño, M. A.; Truhlar, D. G.; Cramer, C. J.; Gagliardi, L. Computational Design of Functionalized Metal-Organic Framework Nodes for Catalysis. ACS Cent. Sci. 2018, 4 (1), 5−19. (24) (a) Teplensky, M. H.; Fantham, M.; Li, P.; Wang, T. C.; Mehta, J. P.; Young, L. J.; Moghadam, P. Z.; Hupp, J. T.; Farha, O. K.; Kaminski, C. F.; Fairen-Jimenez, D. Temperature Treatment of Highly Porous Zirconium-Containing Metal-Organic Frameworks Extends Drug Delivery Release. J. Am. Chem. Soc. 2017, 139, 7522− 7532. (b) Huxford, R. C.; Rocca, J. D.; Lin, W. Metal-Organic Frameworks as Potential Drug Carriers. Curr. Opin. Chem. Biol. 2010, 14, 262−268. (25) (a) Mon, M.; Bruno, R.; Ferrando-Soria, J.; Armentano, D.; Pardo, E. Metal-Organic Framework Technologies for Water Remediation: Towards a Sustainable Ecosystem. J. Mater. Chem. A 2018, 6, 4912−4947. (b) Howarth, A. J.; Katz, M. J.; Wang, T. C.; Platero-Prats, A. E.; Chapman, K. W.; Hupp, J. T.; Farha, O. K. High Efficiency Adsorption and Eemoval of Selenate and Selenite from Water Using Metal-Organic Frameworks. J. Am. Chem. Soc. 2015, 137, 7488−7494. (26) Drout, R. J.; Otake, K.; Howarth, A. J.; Islamoglu, T.; Zhu, L.; Xiao, C.; Wang, S.; Farha, O. K. Efficient Capture of Perrhenate and Pertechnetate by a Mesoporous Zr Metal−Organic Framework and Examination of Anion Binding Motifs. Chem. Mater. 2018, 30, 1277− 1284. (27) Katz, M. J.; Brown, Z. J.; Colon, Y. J.; Siu, P. W.; Scheidt, K. A.; Snurr, R. Q.; Hupp, J. T.; Farha, O. K. A facile synthesis of UiO-66, UiO-67 and their derivatives. Chem. Commun. 2013, 49, 9449−9451. (28) (a) Mondloch, J. E.; Katz, M. J.; Isley, W. C., III; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; Snurr, R. Q.; Cramer, C. J.; Hupp, J. T.; Farha, O. K. Destruction of chemical warfare agents using metal−organic frameworks. Nat. Mater. 2015, 14, 512−516. (b) Islamoglu, T.; Otake, K.I.; Li, P.; Buru, C. T.; Peters, A. W.; Akpinar, I.; Garibay, S. J.; Farha, O. K. Revisiting the structural homogeneity of NU-1000, a Zr-based metal−organic framework. CrystEngComm 2018, 20, 5913−5918. (29) Pang, J.; Yuan, S.; Qin, J.; Liu, C.; Lollar, C.; Wu, M.; Yuan, D.; Zhou, H. C.; Hong, M. Control the Structure of Zr-Tetracarboxylate Frameworks through Steric Tuning. J. Am. Chem. Soc. 2017, 139, 16939−16945. (30) (a) Liu, W.-G.; Truhlar, D. G. Computational Linker Design for Highly Crystalline Metal-Organicf Framework NU-1000. Chem. Mater. 2017, 29, 8073−8081. (b) Goswami, S.; Ray, D.; Otake, K.-i.; Kung, C.-W.; Garibay, S. J.; Islamoglu, T.; Atilgan, A.; Cui, Y.; Cramer, C. J.; Farha, O. K.; Hupp, J. T. A porous, electrically conductive hexa- zirconium(IV) metal−organic framework. Chem. Sci. 2018, 9, 4477−4482. (31) Liu, T.-F.; Vermeulen, N. A.; Howarth, A. J.; Li, P.; Sarjeant, A. A.; Hupp, J. T.; Farha, O. K. Adding to the Arsenal of ZirconiumBased Metal−Organic Frameworks: the Topology as a Platform for

Otagiri, M.; Curry, S. J. Structural Basis of the Drug-binding Specificity of Huma Serum Albumin. J. Mol. Biol. 2005, 353, 38−52. (6) Brettschneider, F.; Tolle, M.; von der Giet, M.; Passlick-Deetjen, J.; Steppan, S.; Peter, M.; Jankowski, V.; Krause, A.; Kuhne, S.; Zidek, W.; Jankowski, J. Removal of protein-bound, hydrophobic uremic toxins by a combined fractionated plasma separation and adsorption technique. Artif. Organs 2013, 37, 409−416. (7) Pavlenko, D.; Giasafaki, D.; Charalambopoulou, G.; van Geffen, E.; Gerritsen, K. G. F.; Steriotis, T.; Stamatialis, D. Carbon Adsorbents With Dual Porosity for Efficient Removal of Uremic Toxins and Cytokines from Human Plasma. Sci. Rep. 2017, 7, 14914. (8) Koubaissy, B.; Toufaily, J.; Yaseen, Z.; Daou, T. J.; Jradi, S.; Hamieh, T. Adsorption of uremic toxins over dealuminated zeolites. Adsorpt. Sci. Technol. 2017, 35, 3−19. (9) Bergé-Lefranc, D.; Pizzala, H.; Paillaud, J. L.; Schäf, O.; Vagner, C.; Boulet, P.; Kuchta, B.; Denoyel, R. Adsorption of small ureic toxin molecules on MEI type zeolites from aqueous solution. Adsorption 2008, 14, 377−387. (10) Wernert, V.; Schaf, O.; Faure, V.; Brunet, P.; Dou, L.; Berland, Y.; Boulet, P.; Kuchta, B.; Denoyel, R. Adsorption of the uremic toxin p-cresol onto hemodialysis membranes and microporous adsorbent zeolite silicalite. J. Biotechnol. 2006, 123, 164−173. (11) Wernert, V.; Schäf, O.; Ghobarkar, H.; Denoyel, R. Adsorption properties of zeolites for artificial kidney applications. Microporous Mesoporous Mater. 2005, 83, 101−113. (12) Tijink, M. S.; Wester, M.; Glorieux, G.; Gerritsen, K. G.; Sun, J.; Swart, P. C.; Borneman, Z.; Wessling, M.; Vanholder, R.; Joles, J. A.; Stamatialis, D. Mixed matrix hollow fiber membranes for removal of protein-bound toxins from human plasma. Biomaterials 2013, 34, 7819−7828. (13) Lu, L.; Yeow, J. T. W. An adsorption study of indoxyl sulfate by zeolites and polyethersulfone-zeolite composite. Mater. Des. 2017, 120, 328−335. (14) Namekawa, K.; Tokoro Schreiber, M.; Aoyagi, T.; Ebara, M. Fabrication of zeolite-polymer composite nanofibers for remobal of uremic toxins from kidney failure patients. Biomater. Sci. 2014, 2, 674−679. (15) Schulman, G.; Agarwal, R.; Acharya, M.; Berl, T.; Blumenthal, S.; Kopyt, N. A multicenter, Randomized, double-Blind, PlaceboControlled, Dose-Ranging Study of AST-120 (Kremezin) in Patients With Moderate to Severe CKD. Am. J. Kidney Dis. 2006, 47 (4), 565− 677. (16) (a) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, C. E.; Sarjeant, A. A.; Snurr, R. Q.; Nguyen, S. T.; Yazaydin, A. O.; Hupp, J. T. Metal-Organic Framework Materials with Ultrahigh Surface Areas: Is the Sky the Limit? J. Am. Chem. Soc. 2012, 134, 15016−15021. (b) Li, P.; Vermeulen, N. A.; Malliakas, C. D.; Gómezgualdrón, D. A.; Howarth, A. J.; Mehdi, B. L.; Dohnalkova, A.; Browning, N. D.; O’Keeffe, M.; Farha, O. K. Bottom-up Construction of a Superstructure in a Porous Uranium-Organic Crystal. Science 2017, 356, 624−627. (17) (a) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705−714. (b) Honicke, I. M.; Senkovska, I.; Bon, V.; Baburin, I. A.; Bonisch, N.; Raschke, S.; Evans, J. D.; Kaskel, S. Balancing Mechanical Stability and Ultrahigh Porosity in Crystalline Framework Materials. Angew. Chem., Int. Ed. 2018, 57, 13780−13783. (18) (a) Yang, X.-C.; Liu, C.; Cao, Y.-M.; Yan, X.-P. Metal-organic framework MIL-100(FE) for artificicial kidney application. RSC Adv. 2014, 4, 40824−40827. (b) Abdelhameed, R. M.; Rehan, M.; Emam, E. H. Figuration of Zr-based MOF@cotton fabric composite for potential kidney application. Carbohydr. Polym. 2018, 195, 460−467. (19) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. (20) (a) Matsuda, R.; Kitaura, R.; Kitagawa, S.; Kubota, Y.; Belosludov, R. V.; Kobayashi, T. C.; Sakamoto, H.; Chiba, T.; Takata, M.; Kawazoe, Y.; Mita, Y. highly controlled acetylene accommodation H

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Article

Journal of the American Chemical Society Solvent-Assisted Metal Incorporation. Eur. J. Inorg. Chem. 2016, 2016, 4349−4352. (32) Furukawa, H.; Gandara, F.; Zhang, Y. B.; Jiang, J.; Queen, W. L.; Hudson, M. R.; Yaghi, O. M. Water Adsorption in Porous MetalOrganic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, 136, 4369−4381. (33) Yabushita, M.; Li, P.; Durkin, K. A.; Kobayashi, H.; Fukuoka, A.; Farha, O. K.; Katz, A. Insights into Supramolecular Sites Responsible for Composite Separation of Biomass-Derived Phenolics and Glucose in Metal-Organic Framework NU-1000. Langmuir 2017, 33, 4129−4137. (34) (a) Sinnokrot, M. O.; Sherrill, C. D. Highly Accurate Coupled Cluster Potential Energy Curves for the Benzene Dimer: Sandwich, TShaped, and Parallel-Displaced Configurations. J. Phys. Chem. A 2004, 108, 10200−10207. (b) Macías, B.; villa, N. V.; Gómez, B.; Borrás, J.; Alzuet, G.; González-Á lvarez, M.; Castiñeiras, A. NDA interaction of new copper(II) complexes with sulfonamides as ligands. J. Inorg. Biochem. 2007, 101, 444−451. (c) Janiak, C. A critical accout on π−π stacking in metal complexes with aromatic nitrogen-containing ligands. J. Chem. Soc., Dalton Trans. 2000, 0, 3885−3896. (35) Langmuir, I. The Constitution and Fundamental Properties of Solids and Liquids. J. Am. Chem. Soc. 1917, 39, 1848−1906. (36) Freundlich, H. Over the Adsorption in Solution. J. Phys. Chem. 1906, 57, 385−470. (37) Deria, P.; Mondloch, J. E.; Tylianakis, E.; Ghosh, P.; Bury, W.; Snurr, R. Q.; Hupp, J. T.; Farha, O. K. Perfluoroalkane Functionalization of NU-1000 via Solvent-Assisted Ligand Incorporation: Synthesis and CO2 Adsorption Studies. J. Am. Chem. Soc. 2013, 135, 16801−16804. (38) Lin, C. J.; Chen, H. H.; Pan, C. F.; Chuang, C. K.; Wang, T. J.; Sun, F. J.; Wu, C. J. p-Cresylsulfate and indoxyl sulfate level at different stages of chronic kidney disease. J. Clin. Lab. Anal. 2011, 25, 191−197. (39) Sheldrick, G. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (40) Sheldrick, G., SHELX2017. Programs for crystal structure determination; Universität Göttingen, Germany,2017. (41) Kabuto, C.; Akine, S.; Nemoto, T.; Kwon, E. Nippon Kessho Gakkaishi 2009, 51, 218−224. (42) Spek, A. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2009, 65, 148−155. (43) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 16, Rev. A.03; Wallingford, CT, 2016.

I

DOI: 10.1021/jacs.8b12525 J. Am. Chem. Soc. XXXX, XXX, XXX−XXX